1: %\documentclass[preprint]{aastex}
2: \documentclass[useAMS,usenatbib]{mn2e}
3: \bibliographystyle{mn2e}
4: \usepackage{epsfig}
5: \usepackage{amsmath}
6: %\usepackage{mn}
7: %\usepackage{mn-nat}
8:
9: \newcommand{\be}{\begin{equation}}
10: \newcommand{\beq}{\begin{equation}}
11: \newcommand{\ba}{\begin{eqnarray}}
12: \newcommand{\ee}{\end{equation}}
13: \newcommand{\eeq}{\end{equation}}
14: \newcommand{\ea}{\end{eqnarray}}
15: \newcommand{\msun}{$M_{\odot}\hspace{1mm}$}
16: \newcommand{\wmap}{{\it WMAP }}
17: \newcommand{\kel}{$^{\circ} K\hspace{1mm}$}
18: \newcommand{\hs}{\hspace{1mm}}
19: \newcommand{\vs}{\vspace{2mm}}
20: \newcommand{\lya}{Ly$\alpha \hspace{1mm}$}
21: \newcommand{\ha}{H$\alpha \hspace{1mm}$}
22: \newcommand{\kms}{km s$^{-1}\hspace{1mm}$}
23: \newcommand{\Mpc}{{\rm Mpc}}
24: \newcommand{\dPsidL}{$\partial \log \Psi/\partial \log L_B$}
25: \newcommand{\apj}{ApJ}
26: \newcommand{\aap}{A\&A}
27: \newcommand{\apjl}{ApJL}
28: \newcommand{\mnras}{MNRAS}
29: \newcommand{\aj}{AJ}
30: \newcommand{\apjs}{ApJS}
31: \newcommand{\nat}{{\it Nature}}
32: \newcommand{\araa}{ARA\&A}
33: \newcommand{\pasj}{PASJ}
34: \newcommand{\llya}{$L_{{\rm Ly}\alpha}$}
35: \newcommand{\physrep}{Physics Reports}
36: % definition to produce a "less than or similar to" symbol
37: \def\lsim{~\rlap{$<$}{\lower 1.0ex\hbox{$\sim$}}}
38:
39: % definition to produce a "greater than or similar to" symbol
40: \def\gsim{~\rlap{$>$}{\lower 1.0ex\hbox{$\sim$}}}
41: %
42:
43: %\title[Ly$\alpha$ Radiation Pressure]{Dynamical Effects
44: %of Ly$\alpha$ Radiation Pressure around Galaxies}
45: \title[Ly$\alpha$ Radiation Pressure]{Ly$\alpha$ Driven Outflows Around Star Forming Galaxies}
46:
47: \author[Mark Dijkstra \& Abraham Loeb]{Mark Dijkstra\thanks{E-mail:mdijkstr@cfa.harvard.edu} \& Abraham Loeb\thanks{E-mail:aloeb@cfa.harvard.edu}\\
48: Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA}
49:
50: \def\LaTeX{L\kern-.36em\raise.3ex\hbox{a}\kern-.15em
51: T\kern-.1667em\lower.7ex\hbox{E}\kern-.125emX}
52:
53: \newtheorem{theorem}{Theorem}[section]
54:
55: %\voffset=-15mm
56:
57: \begin{document}
58:
59: \date{\today}
60: \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2007}
61:
62: \maketitle
63:
64: \label{firstpage}
65: \begin{abstract}
66: We present accurate Monte-Carlo calculations of Ly$\alpha$ radiation
67: pressure in a range of models which represent galaxies during various
68: epochs of our Universe. We show that the radiation force that Ly$\alpha$
69: photons exert on hydrogen gas in the neutral intergalactic medium (IGM),
70: that surrounds minihalos that host the first stars, may exceed gravity by
71: orders of magnitude and drive supersonic winds. Ly$\alpha$ radiation
72: pressure may also dominate over gravity in the neutral IGM that surrounds
73: the HII regions produced by the first galaxies. However, the radiation
74: force is likely too weak to result in supersonic outflows in this
75: case. Furthermore, we show that Ly$\alpha$ radiation pressure may drive
76: outflows in the interstellar medium of star forming galaxies that reach
77: hundreds of km s$^{-1}$. This mechanism could also operate at lower
78: redshifts $z\lsim 6$, and may have already been indirectly detected in the
79: spectral line shape of observed Ly$\alpha$ emission lines. \end{abstract}
80:
81: \begin{keywords}
82: cosmology: theory--galaxies: high redshift--radiation mechanisms: general--radiative transfer--ISM: bubbles
83: \end{keywords}
84:
85: \section{Introduction}
86: \label{sec:intro}
87:
88: HII regions around massive stars convert a significant fraction of the
89: total bolometric luminosity of young galaxies into Ly$\alpha$ line emission
90: \citep{PP67,S03}. This Ly$\alpha$ radiation can exert a large force on
91: surrounding neutral gas, as the Ly$\alpha$ transition has a
92: cross-section that is $\sim 7$ orders of magnitude
93: larger than the Thomson cross-section, when averaged over a frequency band as wide as the resonance frequency itself \citep[e.g.][]{Loeb01}. Not
94: surprisingly, the impact of Ly$\alpha$ radiation pressure on the formation
95: of galaxies has been discussed extensively
96: \citep[e.g][]{Cox85,EF86,Bithell90,Haenelt95,Oh02,McKee07}, but the
97: intricacies of Ly$\alpha$ radiative transfer in 3D complicated an accurate
98: numerical treatment of its dynamical effect on the gas. Nevertheless, an
99: approximate estimate can be obtained from simple energy considerations as
100: shown below.
101:
102: Consider a self-gravitating gas cloud of total (baryons + dark matter) mass $M$ and radius $R$ that
103: contains a central Ly$\alpha$ source. The gravitational binding energy of
104: the baryons inside the cloud, $E_B\sim \Omega_bGM^2/(\Omega_m R)$, can be compared to the total energy in the
105: Ly$\alpha$ radiation field inside the cloud, $E_{\alpha}=L_{\alpha}\times
106: t_{\rm trap}$. Here, $L_{\alpha}$ is the Ly$\alpha$ luminosity of the
107: central source (in erg s$^{-1}$), and $t_{\rm trap}$ is the typical
108: trapping time of Ly$\alpha$ photons in the cloud owing to scattering on
109: hydrogen atoms. The Ly$\alpha$ radiation pressure would unbind the baryonic gas
110: from the cloud if $E_{\alpha}> E_B$, i.e. $L_{\alpha}>\Omega_bGM^2/(\Omega_mRt_{\rm trap})$
111: \citep[e.g.][]{Cox85,Bithell90,Oh02}. In this approach, $t_{\rm trap}$ is
112: one of the key parameters in setting the Ly$\alpha$ radiation
113: pressure. Calculations by \citet{Adams75} imply that $t_{\rm trap}\sim 15
114: t_{\rm light}$ for $3\lsim \log \tau_0 \lsim 5.5$, and $t_{\rm trap}\sim
115: 15(\tau_0/10^{5.5})^{1/3} t_{\rm light}$ otherwise, for a static, uniform,
116: infinite slab of material \citep[also see Fig~1 of][]{Bo79}. Here $\tau_0$
117: is the line center optical depth from the center to the edge of the slab,
118: and $t_{\rm light}$ is the light crossing time if the medium were
119: transparent (i.e. $t_{\rm light}=R/c$ in the case of the cloud described
120: above). Note however, that the precise value of $t_{\rm trap}$ depends on
121: other factors including for example, the gas distribution (clumpiness and
122: geometry), the velocity distribution of the gas, and the dust content of
123: the cloud \citep{Bo79}. The Ly$\alpha$ radiation
124: pressure becomes comparable to gravity when
125: \begin{equation}
126: L_{\alpha,{\rm 41}}=1.0\Big{(}\frac{M}{10^{8}M_{\odot}}\Big{)}^{4/3}\Big{(}\frac{16}{1+z}\Big{)}^2\Big{(}\frac{15t_{\rm light}}{t_{\rm trap}}\Big{)},
127: \label{eq:lcrit}
128: \end{equation}
129: where $L_{\alpha}=L_{\alpha,{\rm 41}}\times 10^{41}$ erg s$^{-1}$, and
130: where we have substituted the virial radius of a galaxy
131: mass $M$, $R_{\rm vir}=0.97$
132: kpc $\times(M/10^8M_{\odot})^{1/3}$$(1+z/16)^{-1}$, for $R$ (Eq.~24 of Barkana \& Loeb
133: 2001). For comparison, a star forming galaxy can generate a Ly$\alpha$
134: luminosity of $L_{\alpha}=(10^{42}-10^{43})\times({\rm
135: SFR}/M_{\odot}\hs{\rm yr}^{-1})$ erg s$^{-1}$, where the precise conversion
136: factor depends on the gas metallicity and the stellar initial mass function
137: \citep[e.g.][]{S03}. Therefore, a star formation rate of merely SFR$\gsim
138: 0.01$--$0.1M_{\odot}$ yr$^{-1}$ is needed to generate a Ly$\alpha$
139: luminosity that is capable of unbinding gas from a halo of mass
140: $10^8-10^9M_{\odot}$.
141:
142: Halos of $\la 10^9M_\odot$ are very common at $z\gsim 6$, and have a
143: sufficiently large reservoir of baryons to sustain the above-mentioned star
144: formation rates for a prolonged time. In this paper we provide a more
145: detailed investigation of the magnitude of Ly$\alpha$ radiation pressure in
146: the environment of high-redshift star forming galaxies.
147: In particular, we use a Ly$\alpha$ Monte-Carlo radiative transfer code
148: \citep{mc} to compute Ly$\alpha$ radiation pressure in a wider range of
149: models.
150: Our treatment of radiative transfer and our focus on the environment of
151: high-redshift star forming galaxies, distinguish this paper from previous
152: work. We will show that the radiation force exerted by Ly$\alpha$ photons
153: on neutral hydrogen gas can exceed the gravitational force that binds the gas
154: to its host galaxy by orders of magnitude, and may drive supersonic
155: outflows of neutral gas both in the intergalactic and the interstellar
156: medium.
157:
158: The outline of this paper is as follows.
159: %
160: In \S~\ref{sec:code} we describe how Ly$\alpha$ radiation pressure is
161: computed in the Monte-Carlo radiative transfer code, and show the tests
162: that are performed to test the accuracy of the code.
163: %
164: In \S~\ref{sec:result}, we present our numerical results. Finally,
165: \S~\ref{sec:conc} summarizes the implications of our work and our main
166: conclusions. The cosmological parameter values used throughout our
167: discussion are
168: $(\Omega_m,\Omega_{\Lambda},\Omega_b,h)=(0.27,0.73,0.042,0.70)$
169: \citep{Komatsu08}.
170:
171:
172: \section{Ly$\alpha$ Radiation Pressure}
173: \label{sec:code}
174:
175: The force $F_{\rm rad}$ experienced by an atom in a direction ${\bf n}$ is
176: related to the flux through a plane normal to ${\bf n}$,
177: \begin{equation}
178: F_{\rm rad}=\frac{4\pi}{c}\int d\nu \hs \sigma({\nu}) H(\nu),
179: \label{eq:flux1}
180: \end{equation}
181: where $\sigma(\nu)$ is the Ly$\alpha$ absorption cross-section at frequency
182: $\nu$. The specific flux is given by $H(\nu)=\frac{1}{2}\int d\mu\hs \mu
183: I(\mu,\nu)$, where $I(\nu,\mu)$ is the specific intensity of the radiation
184: field \citep[see, e.g. Eq. 1.113 in][]{RL79}, and $\mu={\bf n}\cdot {\bf
185: k}$ in which ${\bf k}$ denotes the propagation direction of the radiation
186: (i.e. $\mu=1$ for radiation propagating perpendicular to the plane).
187:
188: The specific intensity obeys the radiative transfer equation, which reads
189: (in spherical coordinates)
190: \begin{equation}
191: \mu \frac{\partial I}{\partial r}+\frac{(1-\mu^2)}{r}\frac{\partial
192: I}{\partial \mu}=\chi_{\nu}(J-I)+S_{\nu}(r),
193: \label{eq:RT}
194: \end{equation}
195: where in this equation $\mu \equiv {\bf r}\cdot {\bf k}/|{\bf r}|$,
196: $J(\nu)=\int d\mu\hs I(\mu,\nu)$ denotes the mean intensity, and $S_{\nu}(r)$ the emission function for newly created photons at frequency $\nu$ and radius $r$ (in photons cm$^{-3}$ s$^{-1}$ sr$^{-1}$ Hz$^{-1}$, see e.g. Loeb \& Rybicki
197: 1999). Furthermore, $\chi_{\nu}=\frac{h_P\nu_{\alpha}}{4\pi}\frac{B_{21}}{\sqrt{\pi}\Delta \nu_{\alpha}}\big{(}3n_1-n_2\big{)}\phi(\nu)$ denotes the opacity at frequency $\nu$ \citep[e.g.][]{RL79}, where $h_{\rm p}$ is Planck's constant, $\nu_\alpha=2.46\times 10^{15}$ Hz is the Ly$\alpha$ frequency, $n_{1(2)}$ is the number density of hydrogen atoms in their electronic ground (first excited) state, $B_{21}$ is the Einstein-B coefficient of the $2\rightarrow 1$ transition, $\phi(\nu)$ is the line profile function (e.g. Rybicki \& Lightman 1979, their Eq. 1.79), and $\Delta \nu_{\alpha}=\frac{v_{\rm th}}{c}$. Here, $v_{th}$ is the thermal velocity of the hydrogen atoms in the gas, given by $v_{th}=\sqrt{2k_B T/m_p}$, where $k_B$ is the Boltzmann constant, $T$ the gas temperature, and $m_p$ the proton mass.
198:
199: Under the assumption that $I(\nu,\mu)$ has only a weak dependence on
200: direction (which is reasonable given that Ly$\alpha$ radiation scatters
201: very frequently), $I(\nu,\mu)$ can be expressed as a first-order Taylor
202: expansion in $\mu$, i.e. $I(\nu,\mu)=a(\nu)+b(\nu)\mu$. In this so-called
203: ``Eddington approximation'', the expression for flux simplifies to
204: \citep[e.g.][ their Eq. 1.118]{RL79}
205: \begin{equation}
206: H(\nu)=\frac{1}{3}\frac{dJ(\nu)}{d\tau}=\frac{c}{12\pi}\frac{du(\nu)}{d\tau},
207: \label{eq:flux2}
208: \end{equation} where we have used the relation, $u=4\pi J/c$, in which $u(\nu)$ is the specific energy density in the radiation field at a frequency $\nu$. Furthermore, we have decoupled the gas' absorption and emission functions from the Ly$\alpha$ radiation field, and assumed that all neutral hydrogen atoms are in their electronic ground state (this assumption is justified in more detail in Appendix~\ref{app:n2}), i.e. $n_2=0$ and $n_1=n_H$. Under this assumption, $d\tau=\chi_{\nu}dr=n_H\sigma(\nu)dr$ with $n_H$ being the number density of neutral hydrogen atoms. Substituting this expression back into Eq.~(\ref{eq:flux1}) yields
209: \begin{equation}
210: F_{\rm rad}=\frac{1}{3n_H}\frac{d}{dr}\int d\nu \hs
211: u(\nu)=\frac{1}{3n_H}\frac{dU}{dr},
212: \label{eq:force}
213: \end{equation}
214: where we defined $U \equiv \int d\nu\hs u(\nu)$. Note that the
215: cross-section does not appear in the final expression for the radiation
216: force\footnote{The right-hand-side of Eq.~(\ref{eq:force}) is analogous to
217: the usual pressure gradient force in fluid-dynamics which is not dependent
218: on the scattering cross-section of the fluid particles.}.
219:
220: \subsection{Implementation in Monte-Carlo Technique}
221: \label{sec:radpresmc}
222:
223: In our Monte-Carlo simulation we sample the gas density and velocity fields
224: with $N_s=5000$ concentric spherical shells. The radius, thickness, and volume of
225: shell $j$ are denoted by $r_j$, $dr_j$, and $V_j$, respectively. We compute
226: the radiation force using two approaches:
227:
228: \begin{itemize}
229:
230: \item In the first approach, we calculate the energy density ($U$ in
231: Eq.~\ref{eq:force}) in the Ly$\alpha$ radiation field as a function of
232: radius: Using the Monte-Carlo simulation we compute the average time that
233: photons spend in shell $j$, which we denote by $\langle t \rangle_j$. The
234: total number of photons that is present in shell $j$ at any given time is
235: then given by $N_{\alpha,j}=\dot{N}_{\alpha}\times \langle t \rangle_j$,
236: where $\dot{N}_{\alpha}$ is the rate at which photons are emitted. This
237: yields the energy density, $U_j=N_{\alpha,j}h\nu_{\alpha}/V_j$. Finally, Eq.~(\ref{eq:force}) is used to compute the radiation force on atoms in shell $j$. Note that estimators of the energy density in -and the momentum transfer by- a radiation field in a more general context is discussed by e.g. \citet{Lucy99} and \citet{Lucy07}.
238:
239: \item In the second approach, we calculate the momentum transfer from a
240: Ly$\alpha$ photon to an atom in each scattering event, $\Delta p=h_{\rm p}
241: ({\bf k}_{\rm in}-{\bf k}_{\rm out})/2\pi$. Here, ${\bf k}_{\rm in}$ and ${\bf k}_{\rm out}$ are the photons
242: wavevectors before and after scattering. We compute the average {\it total
243: momentum transfer} (i.e. summed over all scattering events) per photon in
244: shell $j$, $\langle \Delta \mathcal{P}\rangle_j$, and obtain the total
245: momentum transfer from $\dot{P}_{\alpha,j}=\dot{N}_{\alpha}\times \langle
246: \Delta \mathcal{P}\rangle_j$. The force on an individual atom is obtained
247: by dividing by the total number of hydrogen atoms in shell $j$, i.e
248: $F_j=\dot{P}_{\alpha,j}/(V_j\times n_{H,j})$.
249:
250: \end{itemize}
251:
252: Both methods should give identical results, provided that the Eddington
253: approximation holds.
254:
255: \subsection{Test Case: Sources in a Neutral Comoving IGM}
256: \label{sec:test}
257:
258: We begin by considering a Ly$\alpha$ point source at a redshift $z=10$
259: embedded in a neutral intergalactic medium (IGM) that is expanding with the
260: Hubble flow. The photons scatter and diffuse away from the source while
261: Hubble expansion redshifts the photons away from resonance. In this case, the angle-averaged intensity $J(\nu)$ and its radial dependence can be
262: calculated analytically \citep{LR99}. The availability of analytic
263: expressions for $J(\nu,r)$, and therefore the radiation force $F_{\rm rad}$
264: (through Eq.~\ref{eq:force}), makes this a good test case for our code.
265:
266: In Figure~\ref{fig:uden} we plot the radial dependence of the energy
267: density (in erg cm$^{-3}$) in the Ly$\alpha$ radiation field for a
268: model in which the central source is emitting
269: $\dot{N}_{\alpha,54}\times 10^{54}$ photons s$^{-1}$ (where we have
270: introduced the dimensionless quantity $\dot{N}_{\alpha,54}\equiv
271: (\dot{N}_{\alpha}/10^{54}$ photons s$^{-1}$). This corresponds to a
272: luminosity of $L_{\alpha}=\dot{N}_{\alpha,54} \times 1.6\times
273: 10^{43}$ erg s$^{-1}$, which represents a bright Ly$\alpha$ emitting
274: galaxy \citep[e.g.][]{Ouchi08}. The {\it blue dotted line} shows the
275: energy density if the IGM were fully transparent to Ly$\alpha$
276: radiation. In this hypothetical case all photons stream radially
277: outward, and the energy density is given by $L_{\alpha}/(4\pi r^2
278: c)$. The {\it red dashed line} shows the energy density,
279: $U(r)=\frac{4\pi}{c}\int d\nu\hs J(\nu,r)$, derived from the analytic
280: expression for $J(r,\nu)$ given in Loeb \& Rybicki (1999, their
281: Eq.~21), while the {\it black histogram} shows the energy density
282: extracted from the simulation (\S~\ref{sec:radpresmc}). Clearly, the
283: analytic and Monte-Carlo calculations yield consistent
284: results. Scattering reduces the effective speed at which photons
285: propagate radially outward, which enhances their energy density
286: (especially at small radii) relative to the transparent case. At
287: sufficiently large distances however, the photons have redshifted far
288: enough from resonance that they are propagating almost freely to the
289: observer, and the energy density approaches $L_{\alpha}/4\pi r^2
290: c$. We note that at a sufficiently high value of
291: $\dot{N}_{\alpha,54}$, the fraction of hydrogen atoms that populate
292: the 2p (and 2s) levels is non-negligible and our assumption that (almost) all
293: of the atoms populate their electronic ground state becomes invalid
294: (so that the solution for $U(r)$ in Figure~\ref{fig:uden} breaks
295: down). However, as we show in Appendix~\ref{app:n2}, this only occurs
296: when $\dot{N}_{\alpha,54} \gsim 10^7$, well beyond the regime
297: considered in this paper.
298:
299: \begin{figure}
300: \vbox{\centerline{\epsfig{file=fig1.eps,angle=270,width=8.0truecm}}}
301: \caption[]{Radial profile of the energy density $U(r)$ (erg cm$^{-3}$) in
302: the Ly$\alpha$ radiation field surrounding a central source that is
303: emitting $\dot{N}_{\alpha,54} \times 10^{54}$ photons s$^{-1}$ into an expanding neutral IGM. The {\it blue dotted line} shows $U(r)$ if the IGM were fully transparent. The {\it black solid
304: (red dashed) line} shows $U(r)$ when radiative transfer is included using an
305: analytic (Monte-Carlo) approach (see text). Scattering reduces the speed at
306: which Ly$\alpha$ photons are propagating radially outward, increasing
307: $U(r)$ relative to the transparent case.}
308: \label{fig:uden}
309: \end{figure}
310: \begin{figure}
311: \vbox{\centerline{\epsfig{file=fig2.eps,angle=270,width=8.0truecm}}}
312: \caption[]{The ratio of radiation to gravitational force on a hydrogen atom
313: as a function of physical radius in kpc. To scale out the dependence of
314: this ratio on halo mass, $M_h$, and production rate of Ly$\alpha$ photons
315: by the central source, $\dot{N}_{\alpha}$, the vertical axis is normalized
316: by $M_{h,11}/\dot{N}_{\alpha,54}$ (see text). The {\it black dotted (grey solid) line}
317: was obtained by applying Eq.~(\ref{eq:force}) to the energy density $U(r)$
318: that was obtained by using the analytic (Monte-Carlo) approach (see
319: Fig~\ref{fig:uden}). The {\it black solid line} was obtained by directly computing
320: the momentum transfer rate from photons to atoms in the Monte-Carlo code as
321: outlined in \S~\ref{sec:radpresmc}. The results demonstrate that ({\it i})
322: the radiation force exceeds gravity at $r<10
323: (\dot{N}_{\alpha,54}/M_{h,11})$ kpc, and ({\it ii}) both methods yield
324: consistent results.}
325: \label{fig:force1}
326: \end{figure}
327: %
328: In Figure~\ref{fig:force1} we compare the radiation force to the
329: gravitational force on a single hydrogen atom, $F_{\rm grav}=GM(<r)m_p/r^2$,
330: where $M(<r)$ is the total mass enclosed within a radius $r$). We plot the
331: ratio $F_{\rm rad}/F_{\rm grav}$ scaled by $M=10^{11}M_{\odot}$
332: \footnote{The number density of halos more massive than $10^{11}M_{\odot}$
333: at $z=10$ is $\sim 10^{-7}$ comoving Mpc$^{-3}$, implying that these rare
334: halos are among the most massive ones in existence at that early cosmic
335: time.}. The {\it black dotted line (grey solid histogram)} was calculated by applying
336: Eq.~(\ref{eq:force}) to the energy density $U(r)$ that was obtained by
337: using the analytic (Monte-Carlo) approach (also see
338: Fig~\ref{fig:uden}). For comparison, the {\it black solid histogram} was obtained by
339: directly computing the momentum transfer rate from photons to atoms as
340: outlined in \S~\ref{sec:radpresmc}. Figure~\ref{fig:force1} shows that the
341: radiation force overwhelms gravity at small radii. The energy density
342: scales approximately as $\partial \log U/\partial \log r \sim-2.3$
343: (Fig~\ref{fig:uden}). Therefore, $F_{\rm rad}/F_{\rm grav} \propto r^{-1.3}$ and
344: reaches unity at $r \sim 10$ physical kpc.
345:
346: The radiation force increases linearly with $\dot{N}_{\alpha}$ while the
347: gravitational force scales as $M$. Thus, $F_{\rm rad}/F_{\rm grav}$ scales linearly
348: with the ratio $\mathcal{R}\equiv \dot{N}_{\alpha}/M$. To scale out the
349: dependence on $\mathcal{R}$, the vertical axis shows the quantity
350: $({F_{\rm rad}}/{F_{\rm grav}})\times ({M_{11}}/{\dot{N}_{\alpha,54}})$, where $M_{11}=(M/10^{11} M_{\odot})$. For example, if $M_{11}=0.1$ then
351: Figure~\ref{fig:force1} shows that radiation pressure exceeds gravity
352: out to $r=100$ kpc, well beyond the virial radius of a halo of this mass at $r_{\rm vir}\sim 6.6$ kpc.
353:
354: Most importantly, Figure~\ref{fig:force1} shows that the two approaches
355: used to compute the radiation force in the simulation yield consistent
356: results, with a noticeable deviation only at the largest radii ($r\sim 1$
357: Mpc). At large radii most photons stream outwards radially and the
358: Eddington approximation that was used to derive Eq.~(\ref{eq:flux2})
359: becomes increasingly unreliable.
360:
361: Next, we use the radiative transfer code to explore the magnitude of the
362: Ly$\alpha$ radiation pressure for a range of models which represent an
363: evolutionary sequence of structure formation in the Universe. We focus on
364: the Ly$\alpha$ radiation pressure on gas surrounding ({\it i}) the first
365: stars (\S~\ref{sec:firststar}); ({\it ii}) the first galaxies
366: (\S~\ref{sec:firstgalaxy}); and ({\it iii}) the interstellar medium of
367: galaxies (\S~\ref{sec:wind}).
368:
369: \section{Results}
370: \label{sec:result}
371:
372: \subsection{Case I: A Single Massive Star in a Minihalo}
373: \label{sec:firststar}
374: \begin{figure*}
375: \vbox{\centerline{\epsfig{file=fig3.eps,angle=270,width=15.0truecm}}}
376: \caption[]{The energy density in the Ly$\alpha$ radiation field ({\it
377: left panel}), and the ratio between the radiation and the
378: gravitational forces ({\it right panel}) for a single very massive
379: star ($M_*=100M_{\odot}$) in a minihalo $M_h=2\times 10^6M_{\odot}$
380: . The {\it red solid} line represents a model in which the IGM is
381: at the mean cosmic density and undergoes Hubble expansion right
382: outside the virial radius at $r=0.26$ kpc. The {\it blue dotted line}
383: shows a more realistic model in which the IGM is overdense near the
384: virial radius, and in which the intergalactic gas is gravitationally
385: pulled towards the minihalo (see text). The {\it black dashed line}
386: shows the same model as the {\it red solid line} but with the neutral
387: fraction increasing linearly between $r_{\rm vir}$ and $2r_{\rm
388: vir}$. The star ionizes all the gas out to the virial
389: radius. Ly$\alpha$ photons freely propagate until they reach the edge
390: of the HII region, where they are likely to be scattered back into the
391: ionized minihalo. This yields an almost constant radiation energy
392: density. Once outside the HII region, the radiation force dominates
393: over gravity out to $r=10$ kpc and may accelerate neutral gas outside
394: the HII region to velocities of order $\sim 10$ km s$^{-1}$.}
395: \label{fig:caseI}
396: \end{figure*}
397: Numerical simulations of structure formation suggest the first stars that
398: formed in our Universe were massive ($M_\star \sim 100M_{\odot}$), and
399: formed as single objects in dark matter halos with masses of $M\sim
400: 10^6M_{\odot}$ that collapsed at $z>10$
401: \citep[e.g.][]{Haiman96,Abel02,Yoshida06}. Here, we focus our attention on
402: a star with a mass $M_*=100M_{\odot}$ that formed at $z=15$ in a dark
403: matter mini-halo of mass $M=2\times 10^6M_{\odot}$. The star emits
404: $10^{50}$ ionizing photons per second \citep{S02,Abel07}. We assume that
405: the ionizing flux ionizes all the gas out to the virial radius of the dark
406: matter halo ($r_{\rm vir}=0.26$ kpc) but not beyond that radius
407: \citep{Kita04}. Hence, the IGM gas surrounding this central source is
408: assumed to be neutral ($x_{\rm HI}=1.0$) and cold ($T_{\rm gas}=300$K, which
409: corresponds to the temperature of the neutral IGM at $z=15$ due to X-Ray heating, see e.g. Fig~1 of Pritchard \& Loeb 2008).
410:
411: Recombination following photoionization converts $\sim 68\%$ of all
412: ionizing photons into Ly$\alpha$ photons \citep[][p
413: 387]{Osterbrock89}. Hence, the entire halo is a Ly$\alpha$ source that is
414: surrounded by neutral intergalactic gas. To determine the radial dependence
415: of the Ly$\alpha$ production rate ($S_{\nu}$ in Eq.~\ref{eq:RT}), we need
416: to specify the gas density profile. We assume that the gas distribution
417: inside the dark matter halo is described by an NFW-profile with a
418: concentration parameter $C=5$ and a thermal core\footnote{With this
419: gas density profile, the total recombination rate inside the dark matter
420: halo is $\int_0^{r_{\rm vir}}dr\hs 4\pi r^2n_{\rm H}^2\alpha_{\rm rec}\sim
421: 4.5\times 10^{49}$ s$^{-1}$. The total recombination rate can be increased
422: to balance the photoionization rate by introducing a clumping factor
423: $K\equiv \langle n_H^2 \rangle/\langle n_{\rm H} \rangle^2\sim 2$.} at
424: $r<3r_{\rm vir}/4C$ \citep[see][]{Maller04}. We point out however, that our
425: final results are not sensitive to our choice of $S_{\nu}(r)$.
426:
427: Once $S_{\nu}(r)$ has been determined, we find the radius, $r$, at which a
428: Ly$\alpha$ photon is generated in the Monte-Carlo simulation from the
429: relation
430: \begin{equation}
431: R=\frac{1}{N}\int_0^{r}dr\hs 4\pi r^2 n_{\rm H}^2\alpha_{\rm rec},
432: \label{eq:s}
433: \end{equation}
434: where $R$ is a random number between $0$ and $1$, $N=\int_0^{r_{\rm
435: vir}}dr\hs 4\pi r^2 n_{\rm H}^2\alpha_{\rm rec}$ is the total recombination
436: rate inside the dark matter halo, and $\alpha_{\rm rec}=2.6\times 10^{-13}$
437: cm$^3$ s$^{-1}$ is the case-B recombination coefficient at a temperature
438: $T=10^4$ K \citep[e.g.][]{Hui97}. Once the photon is generated, it scatters
439: through the neutral IGM until it has redshifted far enough from resonance
440: that it can escape to the observer.
441:
442: In the {\it left panel} of Figure~\ref{fig:caseI} we show the energy
443: density (in erg cm$^{-3}$) of the Ly$\alpha$ radiation field as a function
444: of radius. The {\it red solid} line represents a model in which we assumed
445: the IGM to follow the mean density and Hubble expansion right outside the
446: virial radius. The {\it blue dotted line} shows a more realistic model in
447: which the IGM is still overdense near the virial radius, and in which the
448: intergalactic gas is gravitationally pulled towards the minihalo (see
449: Dijkstra et al 2007 for a quantitative description of the density and
450: velocity profiles based on the model of Barkana 2004). The {\it black
451: dashed line} shows the same model as the {\it red solid line} but with the neutral fraction increasing
452: linearly between $r_{\rm vir}$ and $2r_{\rm vir}$. This provides a better
453: representation of the fact that the central population III star emits
454: ionizing photons with energies $\gsim 54$ eV, which can photoionize
455: hydrogen (and helium) atoms that lie deeper in the IGM. The goal of this
456: model is to investigate whether our results depend sensitively on the
457: presence of a sharp boundary between HI and HII.
458:
459: All models show that the radiation energy density within the fully ionized
460: minihalo ($r\lsim r_{\rm vir}=0.26$ kpc) has only a weak dependence on
461: radius, i.e. $d\log U/d\log r \gsim -1$. Naively, this may appear
462: surprising given the fact that within the model, no scattering occurs
463: within the virial radius and one may expect the energy density in the
464: Ly$\alpha$ radiation field to scale as $U \propto r^{-2}$. However, in
465: reality $U$ obtains only a weak radial dependence because the radiation can be
466: scattered back into the ionized minihalo as soon as it 'hits' the wall of
467: neutral IGM gas. Ly$\alpha$ photons are therefore trapped inside the
468: ionized minihalo and their energy density is boosted to a value that is
469: only weakly dependent on radius. On the other hand, for $r\gsim r_{\rm
470: vir}$ we find that $d\log U/d\log r \lsim -2$, which is because Ly$\alpha$
471: photons are trapped more efficiently near the edge of the HII region, while
472: they stream freely outwards at larger radii (as in \S~\ref{sec:test} and
473: Fig~\ref{fig:uden}). Figure~\ref{fig:caseI} shows clearly that the radial
474: dependence of the Ly$\alpha$ energy density is not sensitive to the
475: detailed model assumptions about the gas in the IGM.
476:
477: In the {\it right panel} of Figure~\ref{fig:caseI} we show the ratio
478: between the radiation force (Eq.~\ref{eq:force}) and the gravitational
479: force on a single hydrogen atom), $F_{\rm grav}={GM(<r)m_{\rm
480: p}}/{r^{2}}$, where $M(<r)$ is the total (baryons + dark matter) mass
481: enclosed within a radius $r$. In all models, radiation pressure dominates
482: over gravity by as much as $\gsim 2$ orders of magnitude. The radiation
483: force is largest for the models in which the IGM is assumed to be at
484: mean density, because of the $n_{\rm H}^{-1}$ factor in the
485: equation for the radiation force (Eq.~\ref{eq:force}). Note that the spike
486: near $r\sim 2.6$ kpc for the other two models is due to an artificial
487: discontinuity in the IGM velocity field that exists in this model.
488:
489: \begin{figure*}
490: \vbox{\centerline{\epsfig{file=fig4.eps,angle=270,width=15.0truecm}}}
491: \caption[]{Same as Figure~\ref{fig:caseI}, but for the case of a star
492: forming galaxy ($\dot{M}_*=0.34M_{\odot}$ yr$^{-1}$, see text), surrounded
493: by an HII region with a radius $R_{\rm HII}=50$ kpc ($R_{\rm HII}=20$ kpc)
494: for the {\it solid line} ({\it dotted line}), which is in turn surrounded
495: by a fully neutral intergalactic medium (IGM). For the assumed total halo
496: mass of $M_{\rm tot}=10^9M_{\odot}$, the pressure exerted by the Ly$\alpha$
497: photons is not large enough to exceed gravity. However, radiation pressure
498: wins for $M_{\rm tot}=10^8M_{\odot}$, but even in this case the radiation
499: force is not large enough to produce a significant wind speed in the IGM
500: (see text). }
501: \label{fig:caseII}
502: \end{figure*}
503:
504: Ly$\alpha$ radiation pressure may operate throughout the lifetime of the
505: central star. Over a lifetime of $\sim 2.5$ Myr \citep[see Table 4
506: of][]{S02}, this mechanism is capable of accelerating the gas to velocities
507: of $10$ ($50$) km s$^{-1}$ at $r=0.3$ kpc in the model represented by the
508: {\it blue dotted} ({\it red solid}) {\it line}, and to $4$ ($16$) km
509: s$^{-1}$ at $r=0.4$ kpc (the reason for this large difference is that the edge of the HII region lies at $r=0.26$ kpc. Hence, gas at $r=0.3$ kpc is separated by 0.04 kpc from this edge, while gas at $r=0.4$ kpc is separated by a distance that is 3.5 times larger).
510:
511: Thus, Ly$\alpha$ radiation pressure can accelerate the gas to velocities
512: that exceed the escape velocity from the dark matter halo ($v_{\rm esc}\sim
513: \sqrt{2}v_{\rm circ}\sim 8$ km s$^{-1}$) as well as the sound speed of the
514: intergalactic medium ($c_{\rm s}=2.2(T_{\rm gas}/300\hs{\rm K})^{1/2}$ km
515: s$^{-1}$).
516:
517: Note the as the gas is pushed out and its velocity profile changes, the
518: subsequent radiative transfer is altered. For example, we repeated the
519: radiative transfer calculation for models in which gas at $r_{\rm vir}< r
520: \lsim 2r_{\rm vir}$ was accelerated to velocities in the range $10-20$ km
521: s$^{-1}$ (outward) and found a slightly shallower profile for $U(r)$ which lowered
522: the radiation force by a factor of $\sim 3$. Consequently, the acceleration
523: of the gas decreases with time and the actual velocities reached by the gas
524: are lower than the estimates given above by a factor of a few.
525: Nevertheless, the resulting velocities are still substantial.
526:
527: Our calculations imply that Ly$\alpha$ radiation pressure can affect the
528: gas dynamics in the IGM surrounding minihalos that contain the first
529: stars. The impact of Ly$\alpha$ radiation pressure increases with
530: decreasing density of the surrounding gas in the IGM. In practice, the
531: distribution of the IGM is not spherically symmetric. Instead, the density
532: is expected to vary from sightline to sightline (being large along
533: filaments and small along voids). Our results imply that Ly$\alpha$
534: radiation pressure will be most efficient in 'blowing out' the lower
535: density gas. This conjecture is supported by the tendency of Ly$\alpha$
536: photons to preferentially scatter through the low-density gas; their
537: propagation along the path of least resistance would naturally boost up the
538: Ly$\alpha$ flux there. This effect will be moderated by the tendency of the
539: HII region around the first stars to extend further into the low density
540: gas (in 'butterfly'-like patterns, e.g. Abel et al, 1999).
541:
542: If the central star dies in a supernova explosion, then the resulting
543: violent outflow could blow most of the baryons out from the
544: minihalo. However, stars with masses in the range $30M_{\odot}\lsim
545: M_{*}\lsim140M_{\odot}$ and $M_*\gsim 260M_{\odot}$, are not expected to
546: end their lives in a supernova. Instead, these stars collapse directly to a
547: black hole \citep{HW02} and have weak winds (because of the lack of
548: heavy elements in their atmosphere), so that radiation pressure may
549: be the dominant process that affects their surrounding IGM.
550:
551: In summary, Ly$\alpha$ radiation pressure on the neutral IGM around
552: minihalos in which the first stars form, can exceed gravity by orders of
553: magnitude and launch supersonic winds. Our limited analysis does not allow
554: a detailed discussion on the consequences of these winds. This requires 3D
555: simulations with cosmological initial conditions that capture the full IGM
556: density field around the minihalo and that track the evolution of the
557: shocks that may form in the IGM. Such simulation are numerically
558: challenging as they require self-consistent treatment of gas dynamics and
559: Ly$\alpha$ radiative transfer in a moving inhomogeneous medium.
560:
561: \subsection{Case II: A Young Star Forming Galaxy}
562: \label{sec:firstgalaxy}
563:
564: \begin{figure*}
565: \vbox{\centerline{\epsfig{file=fig5.eps,angle=270,width=15.0truecm}}}
566: \caption[]{Same as in Figure~\ref{fig:caseI} and Figure~\ref{fig:caseII},
567: but for models in which a central Ly$\alpha$ source of luminosity
568: $L_{\alpha}$ is surrounded by a thin ($r_{\rm sh,min}=0.9r_{\rm sh,max}$)
569: spherical shell of HI that is expanding at $v_{\rm sh}=200$ km
570: s$^{-1}$. The {\it blue dotted lines} ({\it red solid lines}) represent a
571: model in which $N_{\rm HI}=10^{21}$ cm$^{-2}$ and $r_{\rm sh,max}=0.1$ kpc
572: ($N_{\rm HI}=10^{19}$ cm$^{-2}$ and $r_{\rm sh,max}=1.0$ kpc). When
573: calculating $F_{\rm grav}$, we assumed the total mass
574: enclosed by the supershell to be $10^8M_{\odot}$ (see text). The results
575: strongly suggest that Ly$\alpha$ radiation pressure may be dynamically
576: important in the interstellar medium of galaxies. The {\it total}
577: radiation force on the shell (obtained as a sum over all atoms) may be
578: computed via $F_{\rm tot}=M_F L_{\alpha}$/c, where $M_F$ may be thought of
579: as a force multiplication factor that depends both on the shell's outflow
580: speed, $v_{\rm exp}$, and its HI column density, $N_{\rm HI}$. The
581: dependence of $M_F$ on these parameters is shown in Fig~\ref{fig:fscat}.}
582: \label{fig:caseIII}
583: \end{figure*}
584:
585: Our second case concerns a young galaxy that is forming multiple stars in a
586: dark matter halo of mass $M=10^9 M_{\odot}$ at $z=10$. We assume that the
587: galaxy is converting a fraction $f_*=10\%$ of its baryons into stars over
588: $\sim 0.1t_{\rm H}$ \citep[][]{Wyithe06}, where $t_{\rm
589: H}=[{2}/{3H(z)}]\sim0.49$ Gyr, is the age of the Universe at $z=10$. This
590: translates to a star formation rate of $\dot{M}_*=0.34M_{\odot}$
591: yr$^{-1}$. For population III stars forming out of pristine gas, the total
592: emission rate of ionizing photons is $\dot{N}_{\rm ion}\sim 3\times
593: 10^{53}$ s$^{-1}$ \citep{S02}\footnote{More precisely, the ionizing photon
594: production rate is $\sim 10^{54}$ s$^{-1}$ $\times({\rm
595: SFR}/M_{\odot}\hs{\rm yr}^{-1})$ in the no-mass-loss model of Schaerer
596: (2002) in which metal-free stars form according to a Salpeter IMF with
597: $M_{\rm low}=1M_{\odot}$ and $M_{\rm high}=500 M_{\odot}$ (his model
598: 'B').}. If $\sim 1\%$ of the ionizing photons escape from the galaxy
599: \citep[][]{Chen07,Gnedin08}, then this translates to a Ly$\alpha$
600: luminosity of $L_{\alpha}=3\times 10^{42}$ erg s$^{-1}$. Furthermore, this
601: galaxy can photoionize a spherical HII region of a radius $R_{\rm HII}\sim
602: 50$ physical kpc. Note however, that other ionizing sources
603: would likely exist within this HII region. Indeed, clusters of sources are
604: thought to determine the growth of ionized bubbles during
605: reionization. This results in a characteristic HII region size that is
606: significantly larger than that produced by single source, especially during
607: the later stages of reionization \citep[e.g.][]{F04,McQuinn07}. In this
608: framework, our model represents a star forming galaxy during the early
609: stages of reionization or alternatively a galaxy that lies $50$ kpc away
610: from the edge of a larger ionized bubble.
611:
612: In this particular case, the majority of all recombination events occur in
613: the central galaxy. Thus, we initiate all Ly$\alpha$ photons at $r=0$ in
614: the Monte-Carlo simulation. We assume that the gas is completely ionized
615: out to $R_{\rm HII}=50$ kpc, beyond which it is neutral. As
616: shown in \S~\ref{sec:firststar}, this abrupt transition in the ionized
617: fraction of H in the gas does not affect our results.
618:
619: The {\it left panel} of Figure~\ref{fig:caseII} shows the energy density
620: (in erg cm$^{-3}$) of the Ly$\alpha$ radiation field as a function of
621: radius. The {\it solid line} represents the model discussed above. A kink
622: in the energy density is seen at the edge of the HII region (see
623: \S~\ref{sec:firststar} for a more detailed discussion of the profile). The
624: {\it dotted line} represents a variant of the model in which we have
625: reduced the size of the HII region to $R_{\rm HII}=20$ kpc.
626:
627: The {\it right panel} of Figure~\ref{fig:caseII} shows the ratio between
628: the radiation and the gravitational forces on a single hydrogen atom. In
629: our fiducial model, the radiation force does not exceed gravity; rather, at
630: the edge of the HII region, gravity is $\sim 3$ times stronger. The
631: radiation force becomes equal to the gravitational force if $R_{\rm
632: HII}=20$ kpc. This requires an extremely low [by a factor $\sim (50/20)^3$]
633: escape fraction of ionizing photons, $f_{\rm esc}\sim 6\times10^{-4}$.
634:
635: Alternatively, radiation pressure {\it is} important when the halo mass of
636: the star forming region is reduced to $10^8M_{\odot}$. Halos of this mass are the the most abundant halos at $z\sim 10$ that are capable of cooling via excitation of atomic hydrogen (i.e. their virial
637: temperature just exceeds $T_{\rm vir}\sim 10^4$ K, e.g. Barkana \& Loeb
638: 2001). The total gas reservoir inside these halos is $M_{\rm
639: b}=\frac{\Omega_{\rm b}}{\Omega_{\rm m}}M_{\rm tot}\sim 1.5\times 10^7
640: M_{\odot}$, and so these halos can sustain a star formation rate of
641: $\dot{M}_{*}=0.3M_{\odot}$ yr$^{-1}$ for up to $\sim 50$ Myr. However, even
642: if the radiation force is allowed to operate for $\sim 50$ Myr, we find
643: that radiation pressure cannot accelerate the gas in the IGM to velocities
644: that exceed $\sim 1$ km s$^{-1}$. We therefore conclude that although
645: Ly$\alpha$ pressure may exceed gravity in the neutral IGM that surrounds
646: HII regions around $M_{\rm tot}=10^8M_{\odot}$ halos, the absolute
647: magnitude of the radiation force is too weak to drive the IGM to supersonic
648: velocities.
649:
650: \subsection{Case III: Ly$\alpha$ Driven Galactic Supershells}
651: \label{sec:wind}
652:
653: In principle, Ly$\alpha$ radiation pressure can be important when neutral
654: gas exists in close proximity to a luminous Ly$\alpha$ source. So far, we
655: focused our attention on HI gas in the IGM. However, neutral gas in the
656: interstellar medium (ISM) of the host galaxy is located closer to the
657: Ly$\alpha$ sources and should be exposed to an even stronger Ly$\alpha$
658: radiation pressure. Indeed, it has been demonstrated
659: \citep[e.g.][]{Ahn02,Verhamme08} that scattering of Ly$\alpha$ photons by
660: neutral hydrogen atoms in a thin (with a thickness much smaller than its radius), outflowing 'supershell' of HI gas surrounding
661: the star forming regions can naturally explain two observed phenomena:
662: ({\it i}) the common shift of the Ly$\alpha$ emission line towards the red
663: relative to metal absorption lines and the host galaxy's systemic redshift determined from other nebular recombination lines \citep[e.g.][]{Pettini01,Shapley03}; and ({\it ii}) the asymmetry of the Ly$\alpha$ line with emission extending well into its red wing \citep[e.g.][]{L95,Tapken07}.
664:
665: The existence of thin, outflowing shells of neutral atomic hydrogen around HII regions is confirmed by
666: HI-observations of our own Milky-Way \citep[][]{Heiles84} and other nearby
667: galaxies \citep[e.g.][]{Ryder95}. The largest of these shells, so-called
668: 'supershells', have radii of $r_{\rm max}\sim 1$ kpc
669: \citep[e.g][]{Ryder95,M02} and HI column densities in the range $N_{\rm
670: HI}\sim 10^{19}$--$10^{21}$ cm$^{-2}$
671: \citep[e.g.][]{L95,Kunth98,Verhamme08}. Supershells are thought to be
672: generated by stellar winds or supernovae explosions which sweep-up gas into
673: a thin expanding neutral shell \citep[see e.g.][for a review]{T88}. The
674: back-scattering mechanism attributes both the redshift and asymmetry of the
675: Ly$\alpha$ line to the Doppler boost that Ly$\alpha$ photons undergo as
676: they scatter off the outflow on the far side of the galaxy back towards the
677: observer \citep[e.g.][]{Lee98,Ahn02,Ahn03,Ahn04,Verhamme06,Verhamme08}. It is interesting to investigate whether Ly$\alpha$ radiation
678: pressure may provide an alternative mechanism that determines the
679: supershell kinematics.
680:
681: In Figure~\ref{fig:caseIII} we show the energy density ({\it left panel})
682: and the Ly$\alpha$ radiation force ({\it right panel}) for two models. Both
683: models assume that: ({\it i}) there is a Ly$\alpha$ source at $r=0$ with a
684: luminosity of $L_{\alpha}=10^{43}$ erg s$^{-1}$; ({\it ii}) the emitted
685: Ly$\alpha$ spectrum prior to scattering has a Gaussian shape as a function
686: of photon frequency with a Doppler velocity width of $\sigma=50$ km
687: s$^{-1}$; ({\it iii}) the spatial width of the supershell is $10\%$ of its
688: radius; and ({\it iv}) the shell has an outflow velocity of $v_{\rm}=200$
689: km s$^{-1}$. The {\it blue dotted (red solid) lines} represent a model in
690: which the supershell has a column density of $N_{\rm HI}=10^{21}$
691: ($N_{\rm HI}=10^{19}$) cm$^{-2}$ and a maximum radius that is $r_{\rm
692: sh}=0.1$ kpc ($r_{\rm sh}=1.0$) kpc. Our calculations assume that there is no neutral gas (or dust) interior to the HI supershell.
693: \begin{figure}
694: \vbox{\centerline{\epsfig{file=fig6.eps,angle=270,width=9.0truecm}}}
695: \caption[]{The multiplication factor $M_F$ provides the total force that
696: Ly$\alpha$ photons exert on a spherical HI shell, $F_{\rm tot}=M_F
697: L_{\alpha}/c$, where $L_{\alpha}$ is the Ly$\alpha$ luminosity of the
698: source in erg s$^{-1}$ and $c$ is the speed of light. The plot shows $M_F$
699: as a function of the expansion velocity of the HI shell, $v_{\rm sh}$, for
700: three values of HI column density, and under the assumption that the HI shell surrounds an empty cavity. The parameter $M_F$ provides a measure
701: of the efficiency by which Ly$\alpha$ photons can be 'trapped' by the shell
702: of HI gas. Thus, $M_F$ increases with increasing $N_{\rm HI}$ and
703: decreasing $v_{\rm sh}$. The {\it dotted horizontal lines} show the values of
704: $M_F$ that have been derived in the past (for a static, uniform,
705: infinite slab of material e.g. Adams 1975).}
706: \label{fig:fscat}
707: \end{figure}
708:
709: The {\it left panel} of Figure~\ref{fig:caseIII} shows that inside the
710: supershell the energy density decreases more gradually than $r^{-2}$
711: because of photon trapping (similarly to the previously discussed cases in
712: \S~\ref{sec:firststar}-\S~\ref{sec:firstgalaxy}). The shell with the larger
713: column of HI is more efficient at trapping the Ly$\alpha$ photons, and thus
714: yields a flatter energy density profile. In both models the energy density
715: drops steeply within the supershell (the energy density decreases as $r^{-2}$ outside the shell, if no scattering occurs here).
716:
717: The {\it right panel} of Figure~\ref{fig:caseIII} shows the ratio between
718: the radiation and gravitational forces. Towards the center of the dark
719: matter halo, baryons dominate the mass density and an evaluation of $F_{\rm
720: grav}$ requires assumptions about the radial distribution of the
721: baryons. For simplicity, we consider a fixed total mass interior to the
722: supershell of $M(<r)=10^8M_{\odot}$, so that $F_{\rm
723: grav}(r)=GM(<r)m_{p}/r^{2}$. Note that any assumed mass profile $M(<r)$
724: will not affect the results as long as $F_{\rm rad}\gg F_{\rm grav}$.
725:
726: We find that the radiation force exceeds gravity in both examples under
727: consideration. For $N_{\rm HI}=10^{21}$ cm$^{-2}$ and $r_{\rm sh}=0.1$ kpc,
728: $F_{\rm rad}\sim 10 F_{\rm grav}$. Thus, radiation pressure would have been
729: important even if we had chosen $M\sim 10^{9}M_{\odot}$.
730: %The virial radius of a dark matter halo of mass $M_{\rm tot}$ that virialized at redshift $z$ is given by $r_{\rm vir}=2$ kpc$(M_{\rm tot}/5\times 10^{7}\hs M_{\odot})^{1/3}([1+z]/6)^{-1}$ \citep[see, e.g. Eq. 24 in][]{BL01}. Even if we assume that a fraction $f_b=\frac{\Omega_{\rm b}}{\Omega_{\rm m}}\sim 0.15$ of the total mass, corresponding to a mass of $\sim 10^7 M_{\odot}$, resides in a sphere of radius $r_{\rm b}=\lambda r_{\rm vir}=0.1$ kpc (where $\lambda\sim 0.05$ is the spin-parameter), it is still not possible for .
731: In the model with $N_{\rm HI}=10^{19}$ cm$^{-2}$ and $r_{\rm sh}=1.0$ kpc,
732: $F_{\rm rad}\sim 10^2 F_{\rm grav}$, and radiation pressure would have been
733: important even if $M\sim 10^{10}M_{\odot}$. Hence, our calculations
734: strongly suggest that Ly$\alpha$ radiation pressure may be dynamically
735: important in the ISM of galaxies.
736:
737: The {\it total} Ly$\alpha$ radiation force is obtained by summing the force
738: over all atoms in the supershell. It is interesting to compare this force
739: to $L_{\alpha}/c$. The latter quantity denotes the total momentum transfer
740: rate (force) from the Ly$\alpha$ radiation field to the supershell under
741: the assumption that each Ly$\alpha$ photon is re-emitted isotropically
742: after entering the shell (including multiple scatterings inside the
743: supershell). In Figure~\ref{fig:fscat} we plot the quantity $M_F$ which is
744: defined as
745: \begin{equation}
746: M_F\equiv \frac{\sum_{\rm atoms}F_{\rm rad}}{L_{\alpha}/c},
747: \end{equation}
748: as a function of the expansion velocity of the shell, $v_{\rm sh}$ for
749: three different values of $N_{\rm HI}$.
750:
751: Figure~\ref{fig:fscat} shows that $M_F$, which can be thought of as a
752: force multiplication factor\footnote{This term derives from the
753: (time-dependent) force-multiplication function $M(t)$ that was
754: introduced by \citet{Castor75}, as $F_{\rm rad}\equiv
755: M(t)({\tau_eL_{\rm bol}}/{c})$. Here, $F_{\rm rad}$ is the total
756: force that radiation exerts on a medium, $\tau_e$ is the total optical
757: depth to electron scattering through this medium. The function $M(t)$
758: arises because of the contribution of numerous metal absorption lines
759: to the medium's opacity, and can be as large as $M_{\rm max}(t)\sim
760: 10^3$ in the atmospheres of O-stars \citep{Castor75}.}, greatly
761: exceeds unity for low shell velocities and large HI column
762: densities. The parameter $M_F$ is related to the mean number of times
763: that a Ly$\alpha$ photon 'bounces' back and forth between opposite
764: sides of the expanding shell. For example, $M_F=1$ when all Ly$\alpha$
765: photons enter the shell, scatter once, and then escape from the shell
766: in no preferred direction. On the other hand, $M_F=3$ when all
767: Ly$\alpha$ photons enter the shell, scatter back towards the opposite
768: direction, and then escape in no preferred direction after scattering
769: in the shell for a second time. A schematic illustration of this
770: argument is provided in Figure~\ref{fig:scheme}. Note that when
771: $M_F=1$ ($M_F=3$), each photon spends on average a timescale of
772: $r_{\rm sh}/c$ ($3r_{\rm sh}/c$) in the bubble enclosed by the
773: shell. In other words, the factor $M_F$ relates to the 'trapping
774: time', $t_{\rm trap}$, that denotes the total time over which
775: Ly$\alpha$ photons are trapped inside the supershell\footnote{This
776: argument ignores the time spent on scattering inside the supershell
777: itself. Photons penetrate on average an optical depth $\tau=1$ into
778: the shell. If this corresponds to a physical distance that is
779: significantly smaller than the thickness of the shell (denoted by
780: $\Delta r_{\rm sh}$), then only a tiny fraction of the photons will
781: diffuse through the shell. Hence, when averaged over these photons,
782: ignoring the time spent inside the supershell itself is
783: justified. Alternatively, photons with a mean free path that is at
784: least comparable to the thickness of the shell, only spend a time
785: $\sim \Delta r_{\rm sh}/c \ll r_{\rm sh}/c$ inside the supershell,
786: which provides a negligible contribution to the trapping time.} (see
787: \S~\ref{sec:intro}) through the relation $M_F=t_{\rm trap}/(r_{\rm
788: sh}/c)$. Indeed, when $v_{\rm sh}\rightarrow 0$ we find that $M_F$
789: reproduces the value $15(\tau_0/10^{5.5})$ (indicated by {\it
790: horizontal dotted lines}) that was found by \citet{Adams75} and
791: \citet{Bo79} reasonably well (keeping in mind that these authors
792: derived their result for a static, uniform, infinite slab of material,
793: and assumed different frequency distributions for the emitted
794: Ly$\alpha$ photons).
795:
796: \begin{figure}
797: \vbox{\centerline{\epsfig{file=fig7.eps,angle=0,width=6.0truecm}}}
798: \caption[]{A schematic illustration of the origin of the force
799: multiplication factor $M_F$ in an expanding supershell. The Ly$\alpha$
800: source is located at the center of the expanding HI supershell. The {\it
801: solid line} represents the trajectories of photons that enter the shell,
802: scatter once, and then escape from the shell isotropically (indicated by
803: the {\it dashed lines}). This corresponds to $M_F=1$. On the other hand,
804: the {\it dot-dashed line} represents the trajectories of photons that enter
805: the shell, are scattered back in the opposite direction, and escape with no
806: preferred direction after scattering in the shell for a second time. This
807: corresponds to a case with $M_F=3$. In general, $M_F=t_{\rm trap}/(r_{\rm
808: sh}/c)$.}
809: \label{fig:scheme}
810: \end{figure}
811: \section{Conclusions}
812: \label{sec:conc}
813:
814: We have applied an existing Monte-Carlo Ly$\alpha$ radiative transfer code
815: \citep[described and tested extensively in][]{mc} to the calculation of the
816: pressure that is exerted by Ly$\alpha$ photons on an optically thick
817: medium. This code enabled us to perform (the first) direct, accurate calculations of Ly$\alpha$ radiation pressure, which distinguishes this work from previous
818: discussions on the importance of Ly$\alpha$ radiation pressure in various
819: astrophysical environments.
820:
821: We have focused on a range of models which represent galaxies at different
822: cosmological epochs. In \S~\ref{sec:firststar} we have shown that the
823: Ly$\alpha$ radiation pressure exerted on the neutral intergalactic medium
824: (IGM) surrounding minihalos ($M_{\rm tot}\sim 10^6 M_{\odot}$) in which the
825: first stars form, can exceed gravity by 2--3 orders of magnitude
826: (Fig~\ref{fig:caseI}), and in principle accelerate the gas in the IGM to
827: tens of km s$^{-1}$. Thus, Ly$\alpha$ radiation pressure can launch
828: supersonic winds in the IGM surrounding the first stars. Our analysis did
829: not allow a detailed study of the consequences of these winds. A
830: comprehensive study would require numerical simulations that capture the
831: full IGM density field around minihalos in 3D and track the evolution of
832: the shocks that may form in the IGM together with the Ly$\alpha$ radiative
833: transfer. In this paper, we have also shown that Ly$\alpha$ radiation
834: pressure is important in the neutral IGM that surrounds the HII regions
835: produced by galaxies with a total halo mass of $M_{\rm tot}=10^8M_{\odot}$
836: (Fig~\ref{fig:caseII}. These are the lowest mass, and hence the most
837: abundant, halos in which gas can cool via atomic line excitation. Here,
838: however, the absolute magnitude of the radiation force is too weak to drive
839: the gas to supersonic velocities.
840:
841: Finally, we have shown in \S~\ref{sec:wind} that the Ly$\alpha$ radiation
842: pressure exerted on neutral gas in the interstellar medium (ISM) of a
843: galaxy can also have strong dynamical consequences. In particular, we have
844: found that the Ly$\alpha$ radiation force exerted on an expanding HI
845: supershell can exceed gravity by orders of magnitude
846: (Fig~\ref{fig:caseIII}), for reasonable assumptions about the
847: gravitational force. It is therefore possible that Ly$\alpha$ radiation
848: pressure plays an important role in determining the kinematics of HI
849: supershells around starburst galaxies. We have demonstrated that the total
850: Ly$\alpha$ radiation force on a spherical HI supershell can be written as
851: $F_{\rm rad}=M_F L_{\alpha}/c$, where the 'force-multiplication factor'
852: $M_F$ relates to the average trapping time of Ly$\alpha$ photons in the
853: neutral medium. The factor $M_F$ can greatly exceed unity, as illustrated
854: by Fig~\ref{fig:fscat}. For comparison, the maximum possible radiation
855: force due to continuum radiation\footnote{
856: %Trapping of continuum photons by free electrons results in a similar force multiplication factor for electron column densities well in excess of $N_e\sim 10^{24}$ cm$^{-2}$. For comparison, HI column densities in supershells are $N_{\rm HI}\lsim 10^{21}$ cm$^{-2}$, and no observational evidence exists that supershells predominantly consists of ionized gas.
857: Continuum radiation may exert a force on steady-state outflows
858: (with a constant mass ejection rate, $\dot{M}$) around late-type stars
859: that may significantly exceed $\dot{M}\Delta v\gg L_{\rm bol}/c$
860: \citep[][]{salpeter74,Ivezic95}. This is not because of 'trapping' of
861: continuum photons, but related to the propagation speed of the photons
862: and the wind. We similarly expect the radiative force of trapped
863: Ly$\alpha$ photons in steady-state outflows to potentially exceed
864: $\dot{M}\Delta v\gg M_FL_{\alpha}/c$ (and as argued in this paper, it
865: is possible that $M_FL_{\alpha}/c>L_{\rm bol}/c$).
866: %This is mainly because the total mechanical momentum change of a shell of material as it traverses a distance $dr$ (in a time $dt=dr/v$) must be compared to the total momentum in the radiation field that was emitted by the star during this time $dt=dr/v$. In some cases this may result in a radiative force as large as $\dot{M}\Delta v\sim 20L_{\rm bol/c}$\citep{ivezic95}.
867: Note though that these steady-state outflows are clearly different
868: from those discussed in \S~\ref{sec:wind}, in which a well defined
869: thin shell HI gas is physically separated from the central Ly$\alpha$
870: source.} is $L_{\rm bol}/c$, in which $L_{\rm bol}$ is the
871: bolometric luminosity of the central galaxy. For a typical star
872: forming galaxy, $L_{\alpha}\sim 0.07L_{\rm bol}$ \citep[e.g.][]{PP67},
873: whereas for a galaxy that contains population III stars,
874: $L_{\alpha}\sim 0.24L_{\rm bol}$ \citep{S03}. Hence, the Ly$\alpha$
875: radiation pressure can dominate the maximum possible continuum
876: radiation pressure if $M_F\gsim 14$ (for a normal stellar population),
877: a threshold which is easily exceeded at large column densities of
878: relatively slow-moving HI shells (see Fig~\ref{fig:fscat}).
879:
880: The possibility that Ly$\alpha$ radiation alone can result in a
881: radiation force that exceeds $L_{\rm bol}/c$ is
882: important. \citet{Murray05} have shown that the total momentum carried
883: by radiation from a star-forming region can exceed the total momentum
884: deposited by supernova explosions in it, and so galactic outflows
885: may be driven predominantly by continuum radiation pressure. We have
886: argued that Ly$\alpha$ radiation pressure may in some cases be even
887: more important than continuum radiation pressure, and thus provide the
888: dominant source of pressure on neutral hydrogen in the ISM.
889:
890: The important implication of our last result is that Ly$\alpha$ radiation
891: pressure may drive outflows of HI gas in the ISM. Observations of local
892: starburst galaxies have shown that the presence of outflowing HI gas may be
893: required to avoid complete destruction of the Ly$\alpha$ radiation by dust
894: and to allow its escape from the host galaxies
895: \citep{Kunth98,Hayes08,Ostlin08,Atek08}. At high redshifts, the Ly$\alpha$
896: emission line of galaxies is often redshifted relative to other nebular
897: recombination lines (such as H$\alpha$) and metal absorption lines
898: \citep[e.g.][]{Pettini01,Shapley03}. Furthermore, the spectral shape of the
899: Ly$\alpha$ emission line is typically asymmetric, with emission extending
900: well into the red wing of the line \citep[e.g.][]{L95}. Both of these
901: observations can be explained simultaneously if the observed Ly$\alpha$
902: photons scatter off neutral hydrogen atoms in an outflowing 'supershell' of
903: HI gas that surrounds the star forming regions
904: \citep{L95,T99,Ahn03,Ahn04,Verhamme06,Verhamme08}. The possibility that
905: Ly$\alpha$ radiation pressure may be important in determining the
906: properties of expanding supershells is exciting, and is discussed in more
907: detail in a companion paper (Dijkstra \& Loeb 2008).
908:
909:
910: {\bf Acknowledgments}
911: This work is supported by in part by NASA grant
912: NNX08AL43G, by FQXi, and by Harvard University funds. We thank Christian Tapken and an anonymous referee for helpful constructive comments.
913:
914:
915: \begin{thebibliography}{xx}
916: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
917: \bibitem[Abel et al.(1999)]{Abel99} Abel, T., Norman, M.~L.,
918: \& Madau, P.\ 1999, \apj, 523, 66
919:
920: \bibitem[Abel et al.(2002)]{Abel02} Abel, T., Bryan, G.~L.,
921: \& Norman, M.~L.\ 2002, Science, 295, 93
922:
923: \bibitem[Abel et al.(2007)]{Abel07} Abel, T., Wise, J.~H.,
924: \& Bryan, G.~L.\ 2007, \apjl, 659, L87
925:
926: \bibitem[Adams(1975)]{Adams75} Adams, T.~F.\ 1975, \apj, 201,
927: 350
928:
929: \bibitem[Ahn \& Lee(2002)]{Ahn02} Ahn, S.-H., \& Lee, H.-W.\ 2002, Journal of Korean Astronomical
930:
931: \bibitem[Ahn et al.(2003)]{Ahn03} Ahn, S.-H., Lee, H.-W., \&
932: Lee, H.~M.\ 2003, \mnras, 340, 863
933:
934: \bibitem[Ahn(2004)]{Ahn04} Ahn, S.-H.\ 2004, \apjl, 601, L25
935:
936: \bibitem[Atek et al.(2008)]{Atek08} Atek, H., Kunth, D.,
937: Hayes, M., Ostlin, G., Mas-Hesse, J.~M.,
938: \& .\ 2008, ArXiv e-prints, 805, arXiv:0805.3501
939:
940: \bibitem[Barkana
941: \& Loeb(2001)]{BL01} Barkana, R., \& Loeb, A.\ 2001, \physrep, 349, 125
942: \bibitem[Barkana(2004)]{Barkana04} Barkana, R.\ 2004, \mnras,
943: 347, 59
944:
945: \bibitem[Bithell(1990)]{Bithell90} Bithell, M.\ 1990, \mnras,
946: 244, 738
947:
948: \bibitem[Bonilha et al.(1979)]{Bo79} Bonilha, J.~R.~M., Ferch, R., Salpeter, E.~E., Slater, G., \& Noerdlinger, P.~D.\ 1979, \apj, 233, 649
949:
950: \bibitem[Castor et al.(1975)]{Castor75} Castor, J.~I., Abbott, D.~C., \& Klein, R.~I., 1975, \apj, 195, 157
951:
952: \bibitem[Chandrasekhar(1945)]{Chandra45} Chandrasekhar, S.\ 1945,
953: \apj, 102, 402
954: %\bibitem[Chiu \& Draine(1998)]{1998astro.ph..3209C} Chiu, W.~A., \& Draine, B.~T.\ 1998, ArXiv Astrophysics e-prints, arXiv:astro-ph/9803209
955:
956: \bibitem[Chen et al.(2007)]{Chen07} Chen, H.-W., Prochaska,
957: J.~X., \& Gnedin, N.~Y.\ 2007, \apjl, 667, L125
958:
959: \bibitem[Cox(1985)]{Cox85} Cox, D.~P.\ 1985, \apj, 288, 465
960:
961: \bibitem[Dennison et al.(2005)]{D05} Dennison, B., Turner,
962: B.~E., \& Minter, A.~H.\ 2005, \apj, 633, 309
963:
964: \bibitem[Dijkstra et al.(2006)]{mc} Dijkstra, M., Haiman,
965: Z., \& Spaans, M.\ 2006, \apj, 649, 14
966:
967: \bibitem[Dijkstra et al.(2007)]{igm} Dijkstra, M., Lidz,
968: A., \& Wyithe, J.~S.~B.\ 2007, \mnras, 377, 1175
969:
970: \bibitem[Dijkstra \& Loeb(2008)]{dl} Dijkstra, M., \& Loeb, A.
971: 2008, submitted to \mnras
972:
973: \bibitem[Dijkstra et al.(2008)]{FS} Dijkstra, M., et al.
974: 2008, accepted to \mnras
975:
976: \bibitem[Elitzur \& Ferland(1986)]{EF86} Elitzur, M., \& Ferland, G.~J.\ 1986, \apj, 305, 35
977:
978: \bibitem[Furlanetto et al.(2004)]{F04} Furlanetto, S.~R.,
979: Zaldarriaga, M., \& Hernquist, L.\ 2004, \apj, 613, 1
980:
981: \bibitem[Gnedin et al.(2008)]{Gnedin08} Gnedin, N.~Y., Kravtsov, A.~V., \& Chen, H.-W.\ 2008, \apj, 672, 765
982:
983: \bibitem[Haehnelt(1995)]{Haenelt95} Haehnelt, M.~G.\ 1995,
984: \mnras, 273, 249
985:
986: \bibitem[Haiman et al.(1996)]{Haiman96} Haiman, Z., Thoul,
987: A.~A., \& Loeb, A.\ 1996, \apj, 464, 523
988:
989: \bibitem[Hayes et al.(2008)]{Hayes08} Hayes, M., Ostlin, G.,
990: Mas-Hesse, J.~M., \& Kunth, D.\ 2008, ArXiv e-prints, 803, arXiv:0803.1176
991:
992: \bibitem[Heger \& Woosley(2002)]{HW02} Heger, A., \& Woosley, S.~E.\ 2002, \apj, 567, 532
993:
994: \bibitem[Heiles(1984)]{Heiles84} Heiles, C.\ 1984, \apjs, 55,
995: 585
996:
997: \bibitem[Hirata(2006)]{Hirata06} Hirata, C.~M.\ 2006, \mnras,
998: 367, 259
999:
1000: \bibitem[Hui \& Gnedin(1997)]{Hui97} Hui, L., \& Gnedin, N.~Y.\ 1997, \mnras, 292, 27
1001:
1002: \bibitem[Ivezic \& Elitzur(1995)]{Ivezic95} Ivezic, Z., \& Elitzur, M.\ 1995, \apj, 445, 415
1003:
1004: \bibitem[Kitayama et al.(2004)]{Kita04} Kitayama, T., Yoshida,
1005: N., Susa, H., \& Umemura, M.\ 2004, \apj, 613, 631
1006:
1007: \bibitem[Komatsu et al.(2008)]{Komatsu08} Komatsu, E., et al.\
1008: 2008, ArXiv e-prints, 803, arXiv:0803.0547
1009:
1010: \bibitem[Kunth et al.(1998)]{Kunth98} Kunth, D., Mas-Hesse, J.~M.,
1011: Terlevich, E., Terlevich, R., Lequeux, J., \& Fall, S.~M.\ 1998, \aap, 334,
1012: 11
1013:
1014: \bibitem[Lee
1015: \& Ahn(1998)]{Lee98} Lee, H.-W., \& Ahn, S.-H.\ 1998, \apjl, 504, L61
1016:
1017: \bibitem[Lequeux et al.(1995)]{L95} Lequeux, J., Kunth, D., Mas-Hesse,
1018: J.~M., \& Sargent, W.~L.~W.\ 1995, \aap, 301, 18
1019:
1020: \bibitem[Loeb \& Rybicki(1999)]{LR99} Loeb, A., \& Rybicki, G.~B.\ 1999,
1021: \apj, 524, 527 (LR99)
1022:
1023: \bibitem[Loeb(2001)]{Loeb01} Loeb, A.\ 2001, \apjl, 555, L1
1024:
1025: \bibitem[Lucy(1999)]{Lucy99} Lucy, L.~B.\ 1999, \aap, 344, 282
1026:
1027: \bibitem[Lucy(2007)]{Lucy07} Lucy, L.~B.\ 2007, \aap, 468, 649
1028:
1029: \bibitem[Maller \& Bullock(2004)]{Maller04} Maller, A.~H., \& Bullock,
1030: J.~S.\ 2004, \mnras, 355, 694
1031:
1032: \bibitem[McKee \& Tan(2008)]{McKee07} McKee, C.~F., \& Tan, J.~C.\ 2008, \apj, 681, 771
1033:
1034: \bibitem[McClure-Griffiths et al.(2002)]{M02}
1035: McClure-Griffiths, N.~M., Dickey, J.~M., Gaensler, B.~M.,
1036: \& Green, A.~J.\ 2002, \apj, 578, 176
1037:
1038: \bibitem[McQuinn et al.(2007)]{McQuinn07} McQuinn, M., Lidz, A., Zahn, O., Dutta, S., Hernquist, L.,
1039: \& Zaldarriaga, M.\ 2007, \mnras, 377, 1043
1040:
1041: \bibitem[Murray et al.(2005)]{Murray05} Murray, N., Quataert,
1042: E., \& Thompson, T.~A.\ 2005, \apj, 618, 569
1043:
1044: \bibitem[Oh \& Haiman(2002)]{Oh02} Oh, S.~P., \& Haiman, Z.\ 2002, \apj, 569, 558
1045:
1046: \bibitem[Osterbrock(1989)]{Osterbrock89} Osterbrock, D.~E.\ 1989, {\it Astrophysics of gaseous nebulae and active galactic nuclei}, University of Minnesota, et al.~Mill Valley, CA, University Science Books.
1047:
1048: \bibitem[Ostlin et al.(2008)]{Ostlin08} Ostlin, G., Hayes, M.,
1049: Kunth, D., Mas-Hesse, J.~M., Leitherer, C., Petrosian, A.,
1050: \& Atek, H.\ 2008, ArXiv e-prints, 803, arXiv:0803.1174
1051:
1052: \bibitem[Ouchi et al.(2008)]{Ouchi08} Ouchi, M., et al.\ 2008,
1053: \apjs, 176, 301
1054:
1055: \bibitem[Partridge \& Peebles(1967)]{PP67} Partridge, R.~B., \& Peebles, P.~J.~E.\ 1967, \apj, 147, 868
1056:
1057: \bibitem[Pettini et al.(2001)]{Pettini01} Pettini, M., Shapley,
1058: A.~E., Steidel, C.~C., Cuby, J.-G., Dickinson, M., Moorwood, A.~F.~M.,
1059: Adelberger, K.~L., \& Giavalisco, M.\ 2001, \apj, 554, 981
1060:
1061: \bibitem[Pritchard \& Furlanetto(2006)]{PF06} Pritchard, J.~R., \& Furlanetto, S.~R.\ 2006, \mnras, 367, 1057
1062:
1063: \bibitem[Pritchard
1064: \& Loeb(2008)]{PL08} Pritchard, J.~R., \& Loeb, A.\ 2008, ArXiv e-prints, 802, arXiv:0802.2102
1065:
1066: \bibitem[Rybicki \& Lightman(1979)]{RL79} Rybicki, G.~B., \& Lightman, A.~P.\ 1979, New York, Wiley-Interscience, 1979.~393 p.,
1067:
1068: \bibitem[Ryder et al.(1995)]{Ryder95} Ryder, S.~D.,
1069: Staveley-Smith, L., Malin, D., \& Walsh, W.\ 1995, \aj, 109, 1592
1070:
1071: \bibitem[Salpeter(1974)]{salpeter74} Salpeter, E.~E.\ 1974, \apj, 193, 585
1072:
1073: \bibitem[Schaerer(2002)]{S02} Schaerer, D.\ 2002, \aap, 382, 28
1074:
1075: \bibitem[Schaerer(2003)]{S03} Schaerer, D.\ 2003, \aap, 397, 527
1076:
1077: \bibitem[Sethi et al.(2007)]{Sethi07} Sethi, S.~K.,
1078: Subrahmanyan, R., \& Roshi, D.~A.\ 2007, \apj, 664, 1
1079:
1080: \bibitem[Shapley et al.(2003)]{Shapley03} Shapley, A.~E.,
1081: Steidel, C.~C., Pettini, M., \& Adelberger, K.~L.\ 2003, \apj, 588, 65
1082:
1083: \bibitem[Shimasaku et al.(2006)]{Shima06} Shimasaku, K., et
1084: al.\ 2006, \pasj, 58, 313
1085:
1086: %\bibitem[Schneider et al.(2002)]{S02} Schneider, R.,
1087: %Ferrara, A., Natarajan, P., \& Omukai, K.\ 2002, \apj, 571, 30
1088:
1089: \bibitem[Tapken et al.(2007)]{Tapken07} Tapken, C., Appenzeller, I., Noll,
1090: S., Richling, S., Heidt, J., Meink{\"o}hn, E., \& Mehlert, D.\ 2007, \aap,
1091: 467, 63
1092:
1093: \bibitem[Tenorio-Tagle \& Bodenheimer(1988)]{T88} Tenorio-Tagle, G., \&
1094: Bodenheimer, P.\ 1988, \araa, 26, 145
1095:
1096: \bibitem[Tenorio-Tagle et al.(1999)]{T99} Tenorio-Tagle,
1097: G., Silich, S.~A., Kunth, D., Terlevich, E.,
1098: \& Terlevich, R.\ 1999, \mnras, 309, 332
1099:
1100: \bibitem[Verhamme et al.(2006)]{Verhamme06} Verhamme, A.,
1101: Schaerer, D., \& Maselli, A.\ 2006, \aap, 460, 397
1102:
1103: \bibitem[Verhamme et al.(2008)]{Verhamme08} Verhamme, A., Schaerer, D.,
1104: Atek, H., \& Tapken, C.\ 2008, ArXiv e-prints, 805, arXiv:0805.3601
1105:
1106: \bibitem[Wyithe \& Loeb(2006)]{Wyithe06} Wyithe, J.~S.~B., \& Loeb, A.\ 2006, \nat, 441, 322
1107:
1108: \bibitem[Yoshida et al.(2006)]{Yoshida06} Yoshida, N., Omukai,
1109: K., Hernquist, L., \& Abel, T.\ 2006, \apj, 652, 6
1110:
1111: \end{thebibliography}
1112:
1113: \appendix
1114:
1115: \section{HI Level Populations}
1116: \label{app:n2}
1117: \subsection{The (De)Excitation Rates of the $2p$ Level}
1118:
1119: Our calculations assumed that all neutral hydrogen atoms populate
1120: their electronic ground state (i.e. the $1s$ level). Below, we explore
1121: the physical conditions under which this assumption holds.
1122:
1123: Processes that populate the $2p$ level include: ({\it i}) collisional
1124: excitation of neutral hydrogen atoms in both the $1s$ and $2s$ states
1125: by electrons and protons, ({\it ii}) recombination into the
1126: $2p$ state following photoionization or collisional ionization, ({\it
1127: iii}) photoexcitation of $2p$ level, ({\it iv}) photoexcitation of
1128: $np$ ($n>3$) levels, followed by a radiative cascade that passes
1129: through the $2p$ state. Processes that {\it de-populate} the $2p$ level
1130: include ({\it v}) collisional de-excitation to by electrons, ({\it vi})
1131: stimulated Ly$\alpha$ emission $2p\rightarrow 1s$ following the
1132: absorption of a Ly$\alpha$ photon, and ({\it vii}) spontaneous
1133: emission of a Ly$\alpha$ photon.
1134:
1135: Quantitatively, the rate at which the population in the $2p$ level is
1136: populated is given by
1137:
1138: \begin{eqnarray}
1139: \frac{dn_{2p}}{dt}=C_{1s2p}n_en_{1s}+ C_{\rm 2s2p}n_pn_{\rm 2s}+0.68n_en_{{\rm HII}}\alpha_{\rm rec,B}+\\ \nonumber
1140: \sum_{\rm n=2}^{\infty}P_{n}n_{1s}f_{\rm n2}-C_{2p1s}n_en_{2p}-C_{\rm 2p2s}n_pn_{2p}-P_{2}n_{2p}-n_{2p}A_{\rm 21},
1141: \label{eq:rate}
1142: \end{eqnarray}
1143: where $C_{lu}$'s denote collisional excitation rate coefficients from
1144: level $l$ to $u$ (in cm$^3$ s$^{-1}$), $n_x$ denote number densities
1145: is species 'x' (i.e $n_e$ denotes the electron number density, while
1146: $n_{\rm HII}$ denotes the number density of HII ions), $P_{n}$ denote
1147: photoexcitation rates to the state $np$ (in s$^{-1}$), and $f_{n2}$
1148: denotes the probability that photoexcitation of the $np$ level results
1149: in a radiative cascade that passes through the $2p$ level.
1150:
1151: In the reminder of this Appendix, we will estimate the
1152: order-of-magnitude of each term.
1153:
1154: \begin{itemize}
1155:
1156: \item The Einstein-A coefficient of the Ly$\alpha$ transition is $A_{21}=6.25\times 10^8$ s$^{-1}$. That is, spontaneous emission of Ly$\alpha$ depopulates the $2p$ state at a rate $A_{21}=6.25\times 10^8$ s$^{-1}$.
1157:
1158: \item Collisional excitation of neutral hydrogen atoms in $1s$ level
1159: by electrons populates the $2p$ level at a rate
1160: $C_{1s2p}n_e=n_e\left[\frac{8.629\times10^{-6}}{T^{1/2}}\right]
1161: \left[\frac{\Omega(1s,2p)}{g_2}\frac{g_{2p}}{g_{1s}}\right]e^{-\chi/kT}~{\rm
1162: s^{-1}}$ \citep[e.g.][]{Osterbrock89}. Here, $T$ denotes the gas
1163: temperature in K, $n_e$ is the electron density in cm$^{-3}$, $g_{1s}=1$
1164: and $g_{2p}=3$ are the statistical weights of the $1s$ and $2p$ levels,
1165: $\Omega(1s,2p)=0.40-0.50$ \citep[5000 K $<$ T $<$ 2$\times 10^4$ K,][]{Osterbrock89}, and $\chi=10.2$ eV is
1166: the energy difference between the $1s$ and $2p$ levels. For
1167: temperatures $T< 2\times 10^4$ K, we find that $C_{1s2p}n_e<10^{-10}n_e$
1168: s$^{-1}$.
1169:
1170: \item The collisional excitation rate of neutral hydrogen atoms in
1171: $2s$ level by protons (which dominate over collisions with electrons
1172: by about a factor of $\sim 10$) populates the $2p$ level at a rate
1173: $C_{2s2p}n_p\sim 2\times 10^{-3} n_p$ s$^{-1}$
1174: \citep[e.g.][]{Osterbrock89}. To assess the term $C_{\rm
1175: 2s2p}n_pn_{\rm 2s}$ requires one to compute $n_{2s}$. The fraction of
1176: atoms in the $2s$ state is determined by rates similar to those
1177: mentioned above, except that the $2s$-state is metastable and its
1178: Einstein coefficient is $A_{2s1s}\approx 8$ s$^{-1}$. The $2s$-state may
1179: therefore be overpopulated relative to the $2p$ state by orders of
1180: magnitude \citep[see e.g.][and references therein]{D05}. In close
1181: proximity to a luminous source, the $2s$ level is populated mostly via
1182: transitions of the form $1s\overset{{\rm Ly}\beta}{\rightarrow}3p
1183: \overset{{\rm H}\alpha}{\rightarrow}2s$ \citep{Sethi07,FS}, and
1184: $n_{2s}\sim 0.12 n_{1s} P_3/A_{2s1s}$, where $P_3$ is the rate at which
1185: Ly$\beta$ photons are scattered and the prefactor $0.12$ denotes the
1186: probability that absorption of the Ly$\beta$ is followed by
1187: re-emission of an H$\alpha$ photon \citep[][]{FS}.
1188:
1189: \item The recombination rate into the $2p$ level is given by
1190: $0.68\alpha_{\rm rec,B}n_en_p=1.8\times 10^{-13}(T_{\rm
1191: gas}/10^4\hs{\rm K})^{-0.7}n_en_p$ cm$^{3}$ s$^{-1}$
1192: \citep[e.g.][]{Hui97}, where $n_p$ is the proton density in cm$^{-3}$.
1193:
1194: \item The rate at which transitions of the form $1s\rightarrow np$
1195: occur by absorbing a photon is given by $P_{n}=4\pi
1196: \int\frac{J(\nu)}{h\nu}\sigma_n(\nu)d\nu$. Assuming for simplicity
1197: that $J(\nu)$ does not vary with frequency, i.e. $J(\nu)=J$, we have
1198: \begin{equation}
1199: P_n=\frac{4\pi J f_n \pi e^2}{h\nu_n m_e c},
1200: \end{equation}
1201: where $f_n$ denotes the oscillator strength of the transition, $e$
1202: ($m_e$) the charge (mass) of the electron, and $h\nu_n$ denotes the
1203: energy difference between the $1s$ and $np$ levels. The oscillator
1204: strength decreases rapidly with increasing $n$ \citep[e.g. chapter
1205: 10.5 of][]{RL79}, and in practice we can safely ignore all terms with
1206: $n>2$. We then need not worry about the factors $f_{n2}$ \citep[which
1207: have been computed by][]{PF06,Hirata06}.
1208:
1209: The Ly$\alpha$ scattering rate is given by
1210: \begin{equation}
1211: P_2=\frac{M_F L_{\alpha} f_{2} \pi e^2}{4\pi r^2 \Delta \nu h\nu_{\alpha} m_e c},
1212: \end{equation}
1213: where we replaced $J$ (in erg s$^{-1}$ Hz$^{-1}$ sr$^{-1}$ cm$^{-2}$)
1214: with $J=M_F\frac{L_\alpha}{16\pi^2 r^2 \Delta \nu}$, in which
1215: $L_{\alpha}$ is the Ly$\alpha$ luminosity of the central source (in
1216: erg s$^{-1}$), and $\Delta \nu$ is the frequency range over which
1217: these Ly$\alpha$ photons have been emitted. The factor $M_F$ takes
1218: into account the fact that resonant scattering traps Ly$\alpha$
1219: photons in an optically thick medium (see
1220: \S~\ref{sec:wind}). Substituting fiducial numbers
1221: \begin{equation}
1222: P_2=2 \times 10^2 \hs {\rm
1223: s}^{-1}\hs\Big{(}\frac{L_{\alpha}}{10^{42}\hs{\rm erg}/{\rm
1224: s}}\Big{)}\Big{(}\frac{10^{-3}\nu_{\alpha}}{\Delta
1225: \nu}\Big{)}\Big{(}\frac{{\rm
1226: pc}}{r}\Big{)}^{2}\Big{(}\frac{M_F}{100}\Big{)}.
1227: \label{eq:lya}
1228: \end{equation}
1229:
1230: The rate at which Ly$\beta$ photons scatter can be related to the Ly$\alpha$ luminosity, $L_{\alpha}$, and the equivalent width (EW) of the line, if one writes the specific intensity $J(\nu_{\beta})$ near the Ly$\beta$ resonance in terms of the Ly$\alpha$ luminosity of the central source as $J(\nu_{\beta})=\frac{\lambda_{\beta}}{\nu_{\beta}}\frac{L_{\alpha}}{{\rm EW}}\frac{1}{16\pi^2r^2}$ (note that we assumed that the specific intensity of the continuum remains constant between $\nu_{\alpha}$ and $\nu_{\beta}$). The rate at which Ly$\beta$ photons scatter can then be written as
1231:
1232: \begin{equation}
1233: P_3=2 \times 10^{-3} \hs {\rm s}^{-1}\hs\Big{(}\frac{L_{\alpha}}{10^{42}\hs{\rm erg}\hs {\rm s}^{-1}}\Big{)} \Big{(}\frac{{\rm EW}}{200\hs \AA}\Big{)}^{-1}\Big{(}\frac{r}{{\rm pc}}\Big{)}^{-2}.
1234: \label{eq:lyb}
1235: \end{equation}
1236: Equation~(\ref{eq:lyb}) illustrates that it is very difficult to bring
1237: the ratio $n_{2s}/n_{1s}\sim 0.12 P_3/(A_{\rm 2s1s})$ to unity.
1238:
1239: \item Collisional deexcitation rates relate to the collisional
1240: excitation rates via $C_{lu}=C_{ul}\frac{g_u}{g_l}e^{-\chi/kT}$
1241: \citep[e.g.][]{Osterbrock89}. Combined with the formulas given above, it is
1242: straightforward to verify that the collisional de-excitation rates are
1243: subdominant relative to the rate at which spontaneous Ly$\alpha$
1244: emission de-populates the $2p$ level.
1245:
1246: \item Lastly, Eq~(\ref{eq:lya}) shows that the stimulated emission rate
1247: is $P_2 \ll A_{21}$.
1248:
1249: \end{itemize}
1250:
1251: \subsection{HI Level Populations in this Paper}
1252:
1253: The maximum number density of hydrogen nuclei in this paper is
1254: encountered in \S~\ref{sec:wind}, for the expanding shell of HI gas
1255: with $N_{\rm HI}=10^{21}$ cm$^{-2}$ and a thickness $dr=0.01$ kpc, in
1256: which $n_{\rm max}\sim 30$ cm$^{-3}$. At these column densities, the
1257: shell self-shields against ionizing radiation, and is likely mostly
1258: neutral. For simplicity, let us assume that $n_e=n_p=n_{\rm HI}=30$
1259: cm$^{-3}$. Under these conditions:
1260:
1261: \begin{itemize}
1262:
1263: \item the collisional excitation rate from $1s\rightarrow 2p$ is
1264: $C_{12} < 3\times 10^{-9}$ s$^{-1}$.
1265:
1266: \item the collisional excitation rate from $2s\rightarrow 2p$ per atom
1267: in the 1s state- is $C_{2s2p}n_{2s}n_{p}/n_{1s}\sim 5\times
1268: 10^{-2}n_{2s}/n_{1s}$ s$^{-1}$$=5\times 10^{-3}P_3/A_{2s1s}$ s$^{-1}$. The maximum
1269: Ly$\alpha$ luminosity considered in this paper is $\sim 10^{43}$ erg
1270: s$^{-1}$. Let us conservatively assume that EW$=20$ \AA (rest-frame),
1271: which corresponds roughly to the detection threshold that exists in
1272: narrow-band surveys \citep[e.g.][]{Shima06}. Using Eq~\ref{eq:lyb}, we
1273: find that $P_{3,{\rm max}} \sim 10^{-5}$ s$^{-1}$ (for $r=0.1$ kpc),
1274: and therefore that $C_{2s2p}n_{2s}n_{p}/n_{1s}\sim 10^{-8}$ s$^{-1}$.
1275:
1276: \item the recombination rate is $5\times 10^{-12}(T_{\rm
1277: gas}/10^4\hs{\rm K})^{-0.7}$ s$^{-1}$.
1278:
1279: \item the maximum Ly$\alpha$ scattering rate is (substituting $r=0.1$
1280: kpc, $M_F=100$, $\Delta \nu=0.001\nu_{\alpha}$ into Eq~\ref{eq:lya})
1281: $P_{2,{\rm max}}=0.2$ s$^{-1}$.
1282:
1283: \end{itemize}
1284:
1285: By comparing these rates to the rate at which spontaneous emission of
1286: Ly$\alpha$ depopulates the $2p$ state, $A_{21}=6.25\times 10^{8}$
1287: s$^{-1}$, we find that all excitation rates are $\geq 9$ orders of
1288: magnitude smaller than the de-excitation rate for the wind models
1289: discussed in \S~\ref{sec:wind}. In equilibrium, hydrogen atoms in
1290: their electronic ground ($1s$) state are therefore $\geq 9$ orders of
1291: magnitude more abundant than hydrogen atoms in their first excited
1292: ($2p$) state. Furthermore, as was mentioned above the ratio of atoms in the 2s and 1s levels is given by $n_{2s}/n_{1s}=0.12 P_{3,{\rm max}}/A_{2s1s}\sim 10^{-7}$. Since the densities, Ly$\alpha$ luminosities, and the Ly$\beta$ scattering rates are lower, the fraction of HI atoms in their first excited states are even smaller in other sections of the paper. In conclusion, for all applications
1293: presented in this paper, no accuracy is lost by assuming that all
1294: hydrogen atoms occupy their electronic ground state.
1295:
1296: Finally, in \ref{sec:test} we computed solutions for the radial
1297: dependence of $U(r)$ (Fig~\ref{fig:uden}). The energy density, $U(r)$,
1298: was quoted to depend linearly on the luminosity of the central
1299: source. This is valid unless ({\it i}) the Ly$\alpha$ scattering rate, $P_2> A_{21}=6.25\times 10^8$ s$^{-1}$, or ({\it ii}) the Ly$\beta$ scattering rate exceeds $P_{\rm 3} \gsim 10^2$ s$^{-1}$. In either case, our
1300: assumption that all hydrogen atoms populate their electronic ground
1301: state breaks down. Condition ({\it i}) translates to $L_{\alpha}\gsim 3\times 10^{48}$ erg s$^{-1}(\Delta \nu/0.001\nu_{\alpha})(r/{\rm pc})^2(100/M_F)$ (Eq.~\ref{eq:lya}), while condition ({\it ii}) translates to $L_{\alpha}\gsim 10^{46}$ erg s$^{-1}({\rm EW}/20\hs{\rm \AA})(r/{\rm
1302: pc})^2$ (Eq.~\ref{eq:lyb}). Substituting $r=0.1$ kpc, $M_F=100$ (Fig~\ref{fig:uden} shows that resonant scattering enhances the energy density by a factor of $\gsim 10$ relative to $L_{\alpha}/4\pi r^2 c$), $\Delta \nu=10^{-3}\nu_{\alpha}$
1303: (thermal broadening alone in $T=10^4$ K gas results in $\Delta v\sim 26$ km s$^{-1}$, which translates to $\Delta \nu=10^{-4}\nu_{\alpha}$), and $EW=20$\AA, condition ({\it i}) translates to $L_{\alpha}\gsim 3\times 10^{52}$ erg s$^{-1}$, while condition ({\it ii}) translates to $L_{\alpha}\gsim 10^{50}$ erg s$^{-1}$
1304: The more conservative condition ({\it ii}), $L_{\alpha}\gsim 10^{50}$ erg s$^{-1}$, translates to $\dot{N}_{54} \gsim 10^7$, well beyond the regime considered in this paper.
1305: \label{lastpage}
1306: \end{document}
1307: